DIRECT MULTISEARCH FOR MULTIOBJECTIVE OPTIMIZATION 1 ...

129 downloads 23797 Views 676KB Size Report
We propose a novel multiobjective derivative-free methodology, calling it .... current iterate (a poll center), the poll step in single objective optimization evaluates.
DIRECT MULTISEARCH FOR MULTIOBJECTIVE OPTIMIZATION ´ A. L. CUSTODIO

∗,

J. F. A. MADEIRA† , A. I. F. VAZ‡ , AND L. N. VICENTE§

Abstract. In practical applications of optimization it is common to have several conflicting objective functions to optimize. Frequently, these functions are subject to noise or can be of blackbox type, preventing the use of derivative-based techniques. We propose a novel multiobjective derivative-free methodology, calling it direct multisearch (DMS), which does not aggregate any of the objective functions. Our framework is inspired by the search/poll paradigm of direct-search methods of directional type and uses the concept of Pareto dominance to maintain a list of nondominated points (from which the new iterates or poll centers are chosen). The aim of our method is to generate as many points in the Pareto front as possible from the polling procedure itself, while keeping the whole framework general enough to accommodate other disseminating strategies, in particular when using the (here also) optional search step. DMS generalizes to multiobjective optimization (MOO) all direct-search methods of directional type. We prove under the common assumptions used in direct search for single objective optimization that at least one limit point of the sequence of iterates generated by DMS lies in (a stationary form of) the Pareto front. However, extensive computational experience has shown that our methodology has an impressive capability of generating the whole Pareto front, even without using a search step. Two by-products of this paper are (i) the development of a collection of test problems for MOO and (ii) the extension of performance and data profiles to MOO, allowing a comparison of several solvers on a large set of test problems, in terms of their efficiency and robustness to determine Pareto fronts. Key words. Multiobjective optimization, derivative-free optimization, direct-search methods, positive spanning sets, Pareto dominance, nonsmooth calculus, performance profiles, data profiles. AMS subject classifications. 90C29, 90C30, 90C56.

1. Introduction. Many optimization problems involve the simultaneous optimization of different objectives or goals, often conflictual. In this paper, we are interested in the development of derivative-free methods (see [9]) for Multiobjective optimization (MOO). Such methods are appropriated when computing the derivatives of the functions involved is expensive, unreliable, or even impossible. Frequently, the term black-box is used to describe objective and/or constraint functions for which, given a point, the value of the function is (hopefully) returned and no further information is provided. The significant increase of computational power and software sophistication observed in the last decades opened the possibility of simulating large and complex systems, leading to the optimization of expensive black-box functions. Such type of black-box functions also appear frequently in MOO problems (see, for instance, [23]). ∗ Department of Mathematics, FCT-UNL, Quinta da Torre, 2829-516 Caparica, Portugal ([email protected]). Support for this author was provided by CMA/FCT/UNL under the grant 2009 ISFL-1-297, and by FCT under the grant PTDC/MAT/098214/2008. † IDMEC-IST, TU-Lisbon, Av. Rovisco Pais, 1040-001 Lisboa, Portugal and ISEL, Rua Conselheiro Em´ıdio Navarro, 1, 1959-007 Lisboa ([email protected]). Support for this author was provided by ISEL, IDMEC-IST, and FCT-POCI 2010. ‡ Department of Production and Systems, University of Minho, Campus de Gualtar, 4710-057, Portugal ([email protected]). Support for this author was provided by Algoritmi Research Center and by FCT under the grant PTDC/MAT/098214/2008. § CMUC, Department of Mathematics, University of Coimbra, 3001-454 Coimbra, Portugal ([email protected]). Support for this author was provided by FCT under the grant PTDC/MAT/098214/2008. Most of this work was developed while this author was visiting the Courant Institute of Mathematical Sciences of New York University under a FLAD scholarship.

1

In the classical literature of MOO, solution techniques are typically classified depending on the moment where the decision maker is able to establish preferences relating the different objectives (see [34]). Solution techniques with a prior articulation of preferences require an aggregation criterion before starting the optimization, combining the different objective functions into a single one. In the context of derivativefree optimization, this approach has been followed in [3, 32]. Different approaches can be considered when aggregating objectives, among which min-max formulations, weighted sums and nonlinear approaches (see, for instance, [42]), and goal programming [28]. In any case, the decision maker must associate weights or/and goals with each objective function. Since the original MOO problem is then reduced to a single objective problem, a typical output will consist of a single nondominated point. If the preferences of the decision maker change, the whole optimization procedure needs to be reapplied. Posteriori articulation of preferences solution techniques circumvent these difficulties, by trying to capture the whole Pareto front for the MOO problem. Weightedsum approaches can also be part of these techniques, considering the weights as parameters and varying them in order to capture the whole Pareto front. However, such methods might be time consuming and might not guarantee an even distribution of points, specially when the Pareto front is nonconvex (see [13]). The normal boundary intersection method [14] was proposed to address these difficulties, but it may provide dominated points as part of the final output. The class of posteriori articulation of preferences techniques also includes heuristics such as genetic algorithms [41] and simulated annealing [43]. The herein proposed algorithmic framework is a member of this latter class of techniques, since it does not aggregate any of the objective functions. Instead, it directly extends, from single to multiobjective optimization, a popular class of directional derivative-free methods, called direct search [9, Chapter 7]. Each iteration of these methods can be organized around a search step and a poll step. Given a current iterate (a poll center), the poll step in single objective optimization evaluates the objective function at some neighbor points defined by a positive spanning set and a step size parameter. We do the same for MOO but change the acceptance criterion of new iterates using Pareto dominance, which then requires the updating of a list of (feasible) nondominated points. At each iteration, polling is performed at a point selected from this list and its success is dictated by changes in the list. Our framework encompasses a search step too, whose main purpose is to further disseminate the search process of all the Pareto front. We coined this new methodology direct multisearch (DMS) — as it reduces to direct search when there is only a single objective function. DMS extends to MOO all types of direct-search methods of directional type such as pattern search and generalized pattern search (GPS) [1, 30], generating set search (GSS) [30], and mesh adaptive direct search (MADS) [2]. Our paper is divided as follows. Section 2 describes the proposed DMS algorithmic framework. An example illustrating how DMS works is described in Section 3. The convergence analysis can be found in Section 4 (and in an Appendix for the more technical details), where we prove, using Clarke’s nonsmooth calculus, that at least a limit point of the sequence of iterates generated by DMS lies in (a stationary form of) the Pareto front. Section 5 of this paper provides information about how our extensive numerical experiments were performed, in particular we describe the set of test problems, the 2

solvers selected for comparison, the metrics used to assess the ability to compute Pareto fronts, and the use of performance and data profiles in MOO. In Section 6 we report a summary of our computational findings, showing the effectiveness and robustness of DMS to compute a relatively accurate approximated Pareto front (even when the initial list of nondominated points is initialized with a singleton and no search step is used). The paper ends with some final comments and discussion of future work in Section 7. In the remaining of the Introduction, we present concepts and terminology from MOO used in our paper (see [36] for a more complete treatment). We pose a constrained nonlinear MOO problem in the form: min s.t.

F (x) ≡ (f1 (x), f2 (x), . . . , fm (x)) x ∈ Ω ⊆ Rn ,



where we consider m (≥ 1) real-extended value objective functions or objective function components fi : Rn → R ∪ {+∞}, i = 1, . . . , m (forming the objective function F (x)), and Ω represents the feasible region. When several objective function components are present, given a point, it may be impossible to find another one which simultaneously improves the value of all the functions at the given one. The concept of Pareto dominance is crucial for comparing any two points, and to describe it we will make use of the strict partial order induced by the cone m Rm + = {z ∈ R : z ≥ 0},

defined by F (x) ≺F F (y) ⇐⇒ F (y) − F (x) ∈ Rm + \ {0}. Given two points x, y in Ω, we say that x ≺ y (x dominates y) when F (x) ≺F F (y). We will also say that a set of points in Ω is nondominated (or indifferent) when no point in the set is dominated by another one in the set. As it is well known, the concept of minimization in single objective optimization does not apply to MOO. In MOO problems it is common to have several conflicting objective functions. Finding a point which corresponds to a minima for all the objectives considered, meaning an ideal point, may be an unrealistic task. The concept of Pareto dominance is used to characterize global and local optimality, by defining a Pareto front or frontier as the set of points in Ω nondominated by any other one in Ω. Definition 1.1. A point x∗ ∈ Ω is said to be a global Pareto minimizer of F in Ω if ∄y ∈ Ω such that y ≺ x∗ . If there exists a neighborhood N (x∗ ) of x∗ such that the previous property holds in Ω ∩ N (x∗ ), then x∗ is called a local Pareto minimizer of F . Rigorously speaking, the Pareto front is the set of global Pareto minimizers. However, the convergence results established for DMS are derived in terms of necessary conditions for local Pareto minimization. 2. Direct multisearch for multiobjective optimization. In derivative-free optimization it is common to use an extreme barrier approach to deal with constraints. We adapt the extreme barrier function to multiobjective optimization (MOO) by setting  F (x) if x ∈ Ω, FΩ (x) = (2.1) (+∞, . . . , +∞)⊤ otherwise. 3

When a point is infeasible, the components of the objective function F are not evaluated, and the values of FΩ are set to +∞. This approach allows to deal with black-box type constraints, where only a yes/no type of answer is returned. We present a general description for direct multisearch (DMS), which encompasses algorithms using different globalization strategies, like those based on integer lattices and only requiring simple decrease of the objective function values for accepting new iterates (see, for example, Generalized Pattern Search [1, 30] and Mesh Adaptive Direct Search [2]), and also algorithms whose globalization strategy imposes a sufficient decrease condition for accepting new iterates (like Generating Set Search methods [30]). Following the MOO terminology, described in the Introduction of the paper, the proposed algorithmic framework keeps a list of previously evaluated feasible nondominated points and corresponding step size parameters. This list plays an important role since it is what is returned to the user at the end of a run and since new iterate points (i.e., poll centers) are chosen from it. Also, as we will see later, success is defined by a change in this list. Thus, we need to introduce the concept of iterate list in addition to the concept of iterate point (used in direct-search methods of directional type for single objective optimization). As also happens for these methods in single objective optimization, each iteration is organized around a search step and a poll step, being the latter one responsible for the convergence results. In DMS, the search step is also optional and used to possibly improve algorithmic performance. After having chosen one of the nondominated points (stored in the current iterate list) as the iterate point (or poll center), each poll step performs a local search around it. In both the search and the poll steps, a temporary list of points is created first, which stores all the points in the current iterate list and all the points evaluated during the course of the step. This temporary list will then be filtered, removing all the dominated points and keeping only the nondominated ones. Note that from (2.1), as we will later see in the description of the algorithm, the infeasible points evaluated during the course of the step are trivially removed. The trial list is then extracted from this filtered list of feasible nondominated points, and must necessarily include (for the purposes of the convergence theory) all the nondominated points which belonged to the iterate list considered at the previous iteration. Different criteria can then be chosen to determine the trial list. A natural possibility is to define the trial list exactly as the filtered one. We will discuss this issue in more detail after the presentation of the algorithmic framework. When the trial list Ltrial is different from the current iterate list Lk , the new iterate list Lk+1 is set to Ltrial (successful search or poll step and iteration). Otherwise, Lk+1 = Lk (unsuccessful poll step and iteration). When using a sufficient decrease condition to achieve global convergence, one makes use of a forcing function ρ : (0, +∞) → (0, +∞), i.e., a continuous and nondecreasing function satisfying ρ(t)/t → 0 when t ↓ 0 (see [30]). Typical examples of forcing functions are ρ(t) = t1+a , for a > 0. To write the algorithm in general terms, we will use ρ¯(·) to either represent the forcing function ρ(·) or the constant, zero function. Let D(L) be the set of points dominated by L and let D(L; a) ⊃ D(L) be the set of points whose distance in the ℓ∞ norm to D(L) is no larger than a > 0. For the purposes of the search step, we say that the point x is nondominated if F (x) ∈ / D(L; ρ¯(α)). When considering the poll step, the point x + αd is nondominated if F (x + αd) ∈ / D(L; ρ¯(αkdk)) (where d is a direction used in polling around x). 4

When ρ¯(·) is a forcing function requiring this improvement or decrease results in the imposition of a sufficient decrease condition. As we will see later in the convergence analysis, the set of directions to be used for polling is not required to positively span Rn (although for coherence with the smooth case we will write it so in the algorithm below), and it is not necessarily drawn from a finite set of directions. In the following description of DMS, the elements of the list are pairs of the form (x; α) but for simplicity we will continue to refer to those elements as points since in the majority of the cases our only interest is to appeal to dominancy or nondominancy in the x part. Algorithm 2.1 (Direct Multisearch for MOO). Initialization Choose x0 ∈ Ω with fi (x0 ) < +∞, ∀i ∈ {1, . . . , m}, α0 > 0, 0 < β1 ≤ β2 < 1, and γ ≥ 1. Let D be a (possibly infinite) set of positive spanning sets. Initialize the list of nondominated points and corresponding step size parameters L0 = {(x0 ; α0 )}. For k = 0, 1, 2, . . . 1. Selection of an iterate point: Order the list Lk in some way (some possibilities are discussed later) and select the first item (x; α) ∈ Lk as the current iterate and step size parameter (thus setting (xk ; αk ) = (x; α)). 2. Search step: Compute a finite set of points {zs }s∈S (in a mesh if ρ¯(·) = 0, see Section A.1) and evaluate FΩ at each element. Set Ladd = {(zs ; αk ), s ∈ S}. Call Lf iltered = filter(Lk,Ladd) to eliminate dominated points from Lk ∪ Ladd , using sufficient decrease to see if points in Ladd are nondominated relatively to Lk . Call Ltrial = select(Lf iltered) to determine Ltrial ⊆ Lf iltered . If Ltrial 6= Lk declare the iteration (and the search step) successful, set Lk+1 = Ltrial , and skip the poll step. 3. Poll step: Choose a positive spanning set Dk from the set D. Evaluate FΩ at the set of poll points Pk = {xk + αk d : d ∈ Dk }. Set Ladd = {(xk + αk d; αk ), d ∈ Dk }. Call Lf iltered = filter(Lk,Ladd) to eliminate dominated points from Lk ∪ Ladd , using sufficient decrease to see if points in Ladd are nondominated relatively to Lk . Call Ltrial = select(Lf iltered) to determine Ltrial ⊆ Lf iltered . If Ltrial 6= Lk declare the iteration (and the poll step) as successful and set Lk+1 = Ltrial . Otherwise, declare the iteration (and the poll step) unsuccessful and set Lk+1 = Lk . 4. Step size parameter update: If the iteration was successful then maintain or increase the corresponding step size parameters: αk,new ∈ [αk , γαk ] and replace all the new points (xk + αk d; αk ) in Lk+1 by (xk + αk d; αk,new ), when success is coming from the poll step, or (zs ; αk ) in Lk+1 by (zs ; αk,new ), when success is coming from the search; replace also (xk ; αk ), if in Lk+1 , by (xk ; αk,new ). Otherwise decrease the step size parameter: αk,new ∈ [β1 αk , β2 αk ] and replace the poll pair (xk ; αk ) in Lk+1 by (xk ; αk,new ). Next we address several issues left open during the discussion and presentation of the DMS framework. List initialization. For simplicity, the algorithmic description presented initia5

lized the list with a single point, but different strategies, considering several feasible previously evaluated points, can be used in this initialization, with the goal of improving the algorithmic performance. In Section 6.1, we suggest and numerically test three possible ways of initializing the list. Note that a list initialization can also be regarded as a search step in iteration 0. Ordering the iterate list. The number of elements stored in the list can vary from one to several, depending on the problem characteristics and also on the criteria implemented to determine the trial list. In a practical implementation, when the iterate list stores several points, it may be crucial to order it before selecting a point for polling, as a way of diversifying the search and explore different regions of Ω. A crude ordering strategy could be, for instance, (i) to always add points to the end of the list and (ii) to move a point already selected as a poll center to the end of the list (doing it at the end of an iteration) for a better dissemination of the search of the Pareto front. In Section 6.3 we will consider an ordering strategy defined by selecting the poll centers using the values of a spread metric. Search step and selection of the iterate point. The search step is optional and, in the case of DMS (m > 1), it might act on the iterate list Lk rather than around an individual point. For consistency with single objective optimization (m = 1), we included the selection of the point iterate before the search step. If the search step is skipped or if it fails, this iterate point will then be the poll center. Another reason for this inclusion is to define a step size parameter for the search step. Polling. As in single objective optimization, one either can have a complete poll step, in which every poll point is evaluated, or an opportunistic poll step, in which the points in the poll set are sampled in a given order and sampling is stopped as soon as some form of improvement is found. In the algorithmic framework presented above, we have used complete polling, which can be a wise choice if the goal is to compute the complete Pareto front. Opportunistic polling may be more suitable to deal with functions of considerably expensive evaluation. In this latter case, in order to improve the algorithmic performance, the poll set should be appropriately ordered before polling [10, 12]. Since the convergence results will rely on the analysis of the algorithmic behavior at unsuccessful iterations, which is identical independently of the polling strategy considered (opportunistic or complete), the results hold for both variants without any further modifications. Filtering dominated points. Note that the filtering process of the dominated points does not require comparisons among all the stored points since the current iterate list Lk is already formed by nondominated points. Instead, only each added point will be compared to the others, and, in particular, (i) if any of the points in the list Lk ∪ Ladd dominates a point in Ladd , this added point will be discarded; (ii) if an added point dominates any of the remaining points in the list Lk ∪ Ladd , all such dominated points will be discarded. An algorithmic description of the procedure used for filtering the dominated points can be found in Figure 2.1. Selecting the trial list. As we have pointed out before, a natural candidate for the new iterate list is Ltrial = Lf iltered , in particular if our goal is to determine as many points in the Pareto front as possible. However, other choices Ltrial ⊂ Lf iltered can be considered. A more restrictive strategy, for instance, is to always consider an iterate list formed by a single point. In such a case, success is achieved if the new iterate point dominates the current one. This type of algorithm fits in our framework since it suffices to initialize the list as a singleton and to only consider in Ltrial the point that dominates the one in Lk when it exists, or the point already 6

Algorithm 2.2: [L3 ]=filter(L1 , L2 ) Set L3 = L1 ∪ L2 for allx ∈ L2 for ally ∈ L3 , y 6= x     if y ≺ x    do   then L3 = L3 \{x}     if x ∈ L3 for ally ∈ L3 , y 6= x do       if x ≺ y          then  L3 = L3 \{y}  do     if y ∈ L2 then           then L2 = L2 \{y} Fig. 2.1. Procedure for filtering the dominated points from L1 ∪ L2 (the set union should not allow element repetition), assuming that L1 is already formed by nondominated points.

Algorithm 2.3: [Ltrial ]=select(Lf iltered ) Set Ltrial = Lf iltered

Algorithm 2.4: [Ltrial ]=select(Lf iltered ) if Lk = {x  k } * Lf iltered Choose a x ∈ Lf iltered which dominated xk then Set Ltrial = {x}  else Set Ltrial = Lk Fig. 2.2. Two procedures for selecting the trial list Ltrial from the list of filtered nondominated points Lf iltered . The list Lk represents the iterate list considered at the current iteration. Note that in both algorithms all the nondominated points in Lk are included in Ltrial , as required for the convergence theory.

in Lk , otherwise. An algorithmic description of these two procedures can be found in Figure 2.2. 3. A worked example. To illustrate how Algorithm 2.1 works, we will now describe in detail its application to problem SP1 [25], defined by: ⊤ min F (x) ≡ (x1 − 1)2 + (x1 − x2 )2 , (x1 − x2 )2 + (x2 − 3)2 s.t. − 1 ≤ x1 ≤ 5, − 1 ≤ x2 ≤ 5. As we will do later in Section 6.1 for the numerical experimentations, we will select, here for the purposes of this example, the trial list from the filtered one as in 7

Algorithm 2.3 (setting Ltrial = Lf iltered ). We will order the list by always adding points to the end of it and by moving a point already selected as a poll center to the end of the list (at the end of an iteration). No search step will be performed. Initialization. Let us set the initial point x0 = (1.5, 1.5), corresponding to (f1 (x0 ), f2 (x0 )) = (0.25, 2.25), and initialize the step size parameter as α0 = 1. The step size will be maintained at successful iterations and halved at unsuccessful ones, which corresponds to setting γ = 1 and β1 = β2 = 12 . Set D = D = [I2 −I2 ], where I2 stands for the identity matrix of dimension 2. Initialize the iterate list of nondominated points as L0 = {(x0 ; 1)}. Iteration 0. The algorithm starts by selecting a point from L0 , in this case the only available, (x0 ; α0 ). Since no search step is performed, the feasible points in the poll set P0 = {(1.5, 1.5) + (1, 0), (1.5, 1.5) + (0, 1), (1.5, 1.5) + (−1, 0), (1.5, 1.5) + (0, −1)} are evaluated (the filled diamonds plotted in Iteration 0 of Figure 3.1 represent the corresponding function values). In this case, all the poll points were feasible, thus Ladd = {((2.5, 1.5); 1), ((1.5, 2.5); 1), ((0.5, 1.5); 1), ((1.5, 0.5); 1)}. The nondominated points are filtered from L0 ∪ Ladd , resulting in Lf iltered = {((1.5, 1.5); 1), ((1.5, 2.5); 1)}. Only one of the evaluated poll points remained unfiltered (the circle in Iteration 0 of Figure 3.1 represents its corresponding function value). According to Algorithm 2.3, Ltrial will coincide with Lf iltered . Since there were changes in L0 , the iteration is declared successful, and L1 = Ltrial = Lf iltered , being the step size maintained. The function values corresponding to the points in L1 are represented by squares in Iteration 0 of Figure 3.1. Note that we move the poll point to the end of the list, yielding the new order L1 = {((1.5, 2.5); 1), ((1.5, 1.5); 1)}.

Iteration 1. At the beginning of the new iteration, the algorithm selects a point from the two stored in L1 . Suppose the point (x1 ; α1 ) = ((1.5, 2.5); 1) was selected. In this case, the poll set P1 = {(2.5, 2.5), (1.5, 3.5), (0.5, 2.5), (1.5, 1.5)} is evaluated (again, the corresponding function values are represented by filled diamonds in Iteration 1 of Figure 3.1). Note that, by coincidence, two of the poll points share the same function values. The list Ladd = {((2.5, 2.5); 1), ((1.5, 3.5); 1), ((0.5, 2.5); 1), ((1.5, 1.5); 1)} is formed and L1 ∪ Ladd is filtered. Again, only one of the poll points was nondominated (the corresponding function value is represented by a circle in Iteration 1 of Figure 3.1). Thus, the iteration was successful, the step size was maintained, and the new list is L2 = Ltrial = Lf iltered = {((1.5, 2.5); 1), ((1.5, 1.5); 1), ((2.5, 2.5); 1)} (the corresponding function values are represented by the squares in Iteration 1 of Figure 3.1). Again, we move the poll point (in this case, ((1.5, 2.5); 1)) to the end of the list. Iteration 2. The next iteration begins by selecting (x2 ; α2 ) = ((1.5, 1.5); 1) from the list L2 (a previous poll center). After evaluating the corresponding poll points (the 8

Iteration0

Iteration1

8

8

7 7

6 6

5

f2

f2

5

4

4 3

3 2

2 1

1

0

0.5

1

1.5

2

2.5

3

0

3.5

0

0.5

1

1.5

f1

2

2.5

3

3.5

4

4.5

f1 Iteration2 8

7

6

f2

5

4

3

2

1

0

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

f1

Fig. 3.1. First three iterations of one instance of Algorithm 2.1, when applied to the MOO problem SP1. The empty diamonds represent the function values corresponding to all the evaluated points since the beginning of the optimization process. The filled diamonds represent the function values corresponding to the poll points evaluated at the current iteration. In circles are represented the nondominated points which were evaluated at the current iteration, and in squares the current iterate list of feasible nondominated points.

filled diamonds), all of them are dominated, thus Ltrial = L2 , the iteration is declared as unsuccessful, the corresponding step size is halved, and L3 = {((1.5, 1.5); 0.5), ((2.5, 2.5); 1), ((1.5, 2.5); 1)} (the corresponding function values are represented by the squares in Iteration 2 of Figure 3.1). In Figure 3.2 we can observe the evolution of the optimization process after 10, 20, and 100 iterations. The number of points in the Pareto front is steadily increasing and, after 100 iterations, the corresponding curve is well defined. 4. Convergence analysis. One of the key ingredients in stating global convergence (i.e., convergence from arbitrary starting points) for direct-search methods of directional type is to establish the existence of a subsequence of step size parameters converging to zero. There are two main strategies which can be used to enforce this property in this class of methods: (i) to ensure that all new iterates lie in an integer lattice when the step size is bounded away from zero or (ii) to impose a sufficient decrease condition in the objective function values when accepting new iterates. To derive this result for direct multisearch (DMS), we need the iterates to lie in a compact set in the former case, and the objective functions must be bounded below in the latter situation. Sm Assumption 4.1. The level set L(x0 ) = i=1 Li (x0 ) is compact, where Li (x0 ) = {x ∈ Ω : fi (x) ≤ fi (x0 )}, i = 1, . . . , m. The objective function components of F are bounded below and above in L(x0 ). Proving that a subsequence of step size parameters converges to zero may require significant detail but the core intuition is relatively simple. We will give this intuition now and relegate the rigorous definitions and proofs to the Appendix of the paper. Note that in both variants (i) and (ii) a successful iteration has necessarily produced at least one new feasible nondominated point. 9

Iteration20 8

7

7

6

6

5

5

f2

f2

Iteration10 8

4

4

3

3

2

2

1

0

1

0

1

2

3

4

5

6

7

0

8

0

1

2

3

f1

4

5

6

7

8

f1 Iteration100 8

7

6

f2

5

4

3

2

1

0

0

1

2

3

4

5

6

7

8

f1

Fig. 3.2. Iterations 10, 20, and 100 of one instance of Algorithm 2.1, when applied to the MOO problem SP1. See the caption of Figure 3.1 for details.

In the first variant (i) one restricts the directions used by the algorithm and the scheme for updating the step size so that all potential iterates lie on an integer lattice when the step size is bounded away from zero, see Figure 4.1. Intuitively, if the step size does not go to zero, points in this integer lattice would be separated by a finite and positive distance, and it would therefore be impossible to fit an infinity of iterates inside a bounded level set. So, the only way to have an infinity of new iterates is for the step size to go to zero. The second variant (ii) is to declare an iteration successful only if it produces a nondominated point that has strictly decreased one of the components of the objective function relatively to at least a point in the list. Recall that ρ(α) is a forcing function, and thus a monotonically increasing function of the step size α. Intuitively, insisting on strict decrease of one of the objective function components via a forcing function will make it harder to have a successful step and therefore will generate more unsuccessful poll steps. As we show in the Appendix, we can prove that there must be, in fact, an infinite number of unsuccessful poll steps. Since each unsuccessful poll step reduces the step size by a factor of β2 < 1, the step size would have to go to zero. Again, fully rigorous versions of these intuitive arguments can be found in the Appendix. 4.1. Refining subsequences and directions. The convergence analysis of direct-search methods of directional type for single objective optimization relies on the analysis of the behavior of the algorithm at limit points of sequences of unsuccessful iterates, denoted by refining subsequences (a concept formalized in [1]). The same applies to DMS. Definition 4.1. A subsequence {xk }k∈K of iterates corresponding to unsuccessful poll steps is said to be a refining subsequence if {αk }k∈K converges to zero. Assumption 4.1, Theorems A.1 or A.2, and the updating strategy of the step size parameter allow us to establish the existence of at least a convergent refining subsequence (see, e.g., [9, Section 7.3]). Theorem 4.2. Let Assumption 4.1 hold. Consider a sequence of iterates gene10

x0

Fig. 4.1. An example of an integer lattice where all potential iterates must lie when the step size is bounded away from zero. The example corresponds to coordinate or compass search where Dk = [In − In ] (In is the identity matrix of order n). The figure depicts a finite portion of that − integer lattice, which is given by {x0 + α0 2r z : z ∈ Zn }, where r − is some negative integer (see [9, Line 5 of Page 130]).

rated by Algorithm 2.1 under the scenarios of either Subsection A.1 (integer lattices) or Subsection A.2 (sufficient decrease). Then there is at least one convergent refining subsequence {xk }k∈K . The first stationarity result in our paper will establish appropriate nonnegativity of generalized directional derivatives (see Definition 4.6) computed along certain limit directions, designated as refining directions (a notion formalized in [2]). Definition 4.3. Let x∗ be the limit point of a convergent refining subsequence. If the limit limk∈K ′ dk /kdk k exists, where K ′ ⊆ K and dk ∈ Dk , and if xk +αk dk ∈ Ω, for sufficiently large k ∈ K ′ , then this limit is said to be a refining direction for x∗ . Note that refining directions exist trivially in the unconstrained case Ω = Rn . 4.2. Tangent cones and generalized derivatives. The main theoretical result of this paper states that a limit point of the sequence of iterates generated by a DMS method is Pareto-Clarke stationary. In this subsection we introduce this definition of stationarity as well as other concepts related to nonsmooth calculus [7], required for the presentation and analysis of the DMS framework. In single objective constrained optimization, a critical point has the property that if one moves slightly away from it in any ‘feasible direction’, the objective function does not improve. For multiobjective constrained optimization, the notion of critical point changes somewhat. Essentially, a critical point will be a point on the local Pareto front (see Definition 1.1). As a result, it will have the property that moving slightly away from this point in any ‘feasible direction’ will not yield a better, dominating point. This, in turn, means that as one moves away in a ‘feasible direction’, at least one of the multiple objectives gets worse. We can formalize these intuitive ideas using the concept of Clarke tangent vector for the notion of ‘feasible direction’ and the concept of Pareto-Clarke critical point for the notion of a ‘critical point on the Pareto front’. The definitions of these quantities are as follows. We start by defining the Clarke tangent cone, which we will use to state ParetoClarke first-order stationarity. The Clarke tangent cone is a generalization of the 11

commonly used tangent cone in Nonlinear Programming (NLP) (see, e.g., [38, Definition 12.2 and Figure 12.8]). Such generalization is convenient for our analysis, but should not confuse a reader used to the basic definition of tangent cones in NLP. The definition and notation are taken from [2]. Definition 4.4. A vector d ∈ Rn is said to be a Clarke tangent vector to the set Ω ⊆ Rn at the point x in the closure of Ω if for every sequence {yk } of elements of Ω that converges to x and for every sequence of positive real numbers {tk } converging to zero, there exists a sequence of vectors {wk } converging to d such that yk + tk wk ∈ Ω. The set TΩCl (x) of all Clarke tangent vectors to Ω at x is called the Clarke tangent cone to Ω at x. We will also need the definition of hypertangent cone since it is strongly related to the type of iterates generated by a direct-search method of directional type. The hypertangent cone is the interior of the Clarke tangent cone (when such interior is nonempty). Again we will follow the notation in [2]. Definition 4.5. A vector d ∈ Rn is said to be a hypertangent vector to the set Ω ⊆ Rn at the point x in Ω if there exists a scalar ǫ > 0 such that y + tw ∈ Ω,

∀y ∈ Ω ∩ B(x; ǫ),

w ∈ B(d; ǫ),

and

0 < t < ǫ.

The set of all hypertangent vectors to Ω at x is called the hypertangent cone to Ω at x and is denoted by TΩH (x). Note that the Clarke tangent cone is the closure of the hypertangent one. If we assume that F (x) is Lipschitz continuous near x (meaning that each fi (x), i = 1, . . . , m, is Lipschitz continuous in a neighborhood of x), we can define the Clarke-Jahn generalized derivatives of the individual functions along directions d in the hypertangent cone to Ω at x, fi◦ (x; d) =

lim sup x′ → x, x′ ∈ Ω t ↓ 0, x′ + td ∈ Ω

fi (x′ + td) − fi (x′ ) , i = 1, . . . , m. t

(4.1)

These derivatives are essentially the Clarke generalized directional derivatives [7], extended by Jahn [27] to the constrained setting. The Clarke-Jahn generalized derivatives along directions v in the tangent cone to Ω at x, are computed by taking a limit, i.e., fi◦ (x; v) = limd∈TΩH (x),d→v fi◦ (x; d), for i = 1, . . . , m (see [2]). We are now able to introduce the definition of Pareto-Clarke stationarity which will play a key role in our paper. Definition 4.6. Let F be Lipschitz continuous near a point x∗ ∈ Ω. We say that x∗ is a Pareto-Clarke critical point of F in Ω if, for all directions d ∈ TΩCl (x∗ ), there exists a j = j(d) ∈ {1, . . . , m} such that fj◦ (x∗ ; d) ≥ 0. Definition 4.6 says essentially that there is no direction in the tangent cone that is descent for all the objective functions. If a point is a Pareto minimizer (local or global), then it is necessarily a Pareto-Clarke critical point. By assuming strict differentiability for each component of the objective function at x∗ (meaning that the corresponding Clarke generalized gradient is a singleton), the previous definition of Pareto-Clarke stationarity can be restated using the gradient vectors. Definition 4.7. Let F be strictly differentiable at a point x∗ ∈ Ω. We say that x∗ is a Pareto-Clarke-KKT critical point of F in Ω if, for all directions d ∈ TΩCl (x∗ ), there exists a j = j(d) ∈ {1, . . . , m} such that ∇fj (x∗ )⊤ d ≥ 0. 12

4.3. Convergence results. We are now in a position to state the main convergence result of our paper. Recall that an unsuccessful poll step means that there is no improving (or nondominating) point in the frame or stencil formed by the poll points. If the step size is large, this does not preclude the possibility of a nearby improving point. However, as the step size approaches zero, the poll points allow us to recover the local sensitivities and this, together with some assumption of smoothness, imply that there is no locally improving ‘feasible direction’. Theorem 4.8. Consider a refining subsequence {xk }k∈K converging to x∗ ∈ Ω and a refining direction d for x∗ in TΩH (x∗ ). Assume that F is Lipschitz continuous near x∗ . Then, there exists a j = j(d) ∈ {1, . . . , m} such that fj◦ (x∗ ; d) ≥ 0. Proof. Let {xk }k∈K be a refining subsequence converging to x∗ ∈ Ω and d = limk∈K ′′ dk /kdk k ∈ TΩH (x∗ ) a refining direction for x∗ , with dk ∈ Dk and xk +αk dk ∈ Ω for all k ∈ K ′′ ⊆ K. For j ∈ {1, . . . , m} we have fj◦ (x∗ ; d) =

lim sup x′ → x∗ , x′ ∈ Ω t ↓ 0, x′ + td ∈ Ω

fj (x′ + td) − fj (x′ ) t

fj (xk + αk kdk k(dk /kdk k)) − fj (xk ) − rk αk kdk k k∈K ′′ fj (xk + αk dk ) − fj (xk ) + ρ¯(αk kdk k) ρ¯(αk kdk k) = lim sup − − rk αk kdk k αk kdk k k∈K ′′ fj (xk + αk dk ) − fj (xk ) + ρ¯(αk kdk k) ≥ lim sup . αk kdk k k∈K ′′ ≥ lim sup

The first inequality follows from {xk }k∈K ′′ being a feasible refining subsequence and the fact that xk + αk dk is feasible for k ∈ K ′′ . The term rk is bounded above by ν||d− dk /kdk kk, where ν is the Lipschitz constant of F near x∗ . Note, also, that the limit limk∈K ′′ ρ¯(αk kdk k)/(αk kdk k) is 0 for both globalization strategies (Subsections A.1 and A.2). In the case of using integer lattices (Subsection A.1), one uses ρ¯(·) = 0. When imposing sufficient decrease (Subsection A.2), this limit follows from the properties of the forcing function and Assumption A.5. Since {xk }k∈K is a refining subsequence, for each k ∈ K ′′ , xk + αk dk is not nondominated relatively to Lk . Thus, for each k ∈ K ′′ it is possible to find j(k) ∈ {1, . . . , m} such that fj(k) (xk + αk dk ) − fj(k) (xk ) + ρ¯(αk kdk k) ≥ 0. Since the number of objective functions components is finite, there must exists one, say j = j(d), for which there is an infinite set of indices K ′′′ ⊆ K ′′ such that ◦ fj(d) (x∗ ; d) ≥ lim sup k∈K ′′′

fj(d) (xk + αk dk ) − fj(d) (xk ) + ρ¯(αk kdk k) ≥ 0. αk kdk k

If we assume strict differentiability of F at the point x∗ , the conclusion of the above result will be ∇fj (x∗ )⊤ d ≥ 0. Convergence for a Pareto-Clarke critical point (see Definition 4.6) or a ParetoClarke-KKT critical point (see Definition 4.7) can be established by imposing density in the unit sphere of the set of refining directions associated with x∗ . We note that this assumption is stronger than just considering that the normalized set of directions D is dense in the unit sphere. 13

Theorem 4.9. Consider a refining subsequence {xk }k∈K converging to x∗ ∈ Ω. Assume that F is Lipschitz continuous near x∗ and TΩH (x∗ ) 6= ∅. If the set of refining directions for x∗ is dense in TΩCl (x∗ ), then x∗ is a Pareto-Clarke critical point. If, in addition, F is strictly differentiable at x∗ , then this point is a Pareto-ClarkeKKT critical point. Proof. Given any direction v in the Clarke tangent cone, one has that fj◦ (x∗ ; v) =

lim fj◦ (x∗ ; d), d→v d ∈ TΩH (x∗ )

for all j ∈ {1, . . . , m} (see [2]). Since the number of objective functions is finite, and from the previous theorem, there must exist a sequence of directions {dw }w∈W in TΩH (x∗ ), converging to v such that fj◦ (x∗ ; dw ) ≥ 0 for all directions dw in that sequence and for some j = j(v) ∈ {1, . . . , m}. The first statement of the theorem follows by taking limits of the Clarke generalized derivatives in this sequence (and the second one results trivially). Note that the assumption of density of the set of refining directions in the unit sphere is not required only because of the presence of constraints. In fact, it is also necessary even without constraints because one can easily present examples where the cone of directions simultaneous descent for all objective functions can be as narrow as one would like. In the following corollary, we state the previous results for the particular case of single objective optimization, where the number of the objective function components equals one. Corollary 4.10. Let m = 1 and F = (f1 ) = f . Under the conditions of Theorem 4.8, if d ∈ TΩH (x∗ ) is a refining direction for x∗ , then f ◦ (x∗ ; d) ≥ 0. Under the conditions of Theorem 4.9, the point x∗ is a Clarke critical point, i.e., f ◦ (x∗ ; v) ≥ 0, ∀v ∈ TΩCl (x∗ ). If, additionally, we require the inclusion of all the nondominated points in the iterate list, and if it is finite the number of iterations for which the cardinality of the iterate list exceeds one, we can establish first-order convergence for an ideal point. Corollary 4.11. Consider the algorithmic variant where Ltrial = Lf iltered in all iterations (Algorithm 2.3). Assume that is finite the number of iterations for which the cardinality of {Lk }k∈K exceeds one. Under the conditions of Theorem 4.8, if d ∈ TΩH (x∗ ) is a refining direction for x∗ , we have, for all j ∈ {1, . . . , m}, fj◦ (x∗ ; d) ≥ 0. Under the conditions of Theorem 4.9, the point x∗ is an ideal point, i.e., fj◦ (x∗ ; v) ≥ 0,

∀j ∈ {1, . . . , m}, ∀v ∈ TΩCl (x∗ ).

Proof. Let us recall the proof of Theorem 4.8 until its last paragraph. Now, by assumption, it is possible to consider an infinite subset of indices K ′′′ ⊆ K ′′ such that |Lk | = 1, for each k ∈ K ′′′ . The selection criterion for the iterate list ensures that for each k ∈ K ′′′ , xk + αk dk is dominated by xk and it follows trivially that fj◦ (x∗ ; d) ≥ 0 for all j ∈ {1, . . . , m}. The proof of the second assertion follows the same type of arguments of the proof of Theorem 4.9. 5. Test problems, solvers, metrics, and profiles. 14

5.1. Test problems. We have collected 100 multiobjective optimization (MOO) problems reported in the literature involving only simple bounds constraints, i.e., problems for which Ω = [ℓ, u] with ℓ, u ∈ Rn and ℓ < u. All test problems were modeled by us in AMPL (A Modeling Language for Mathematical Programming) [22] and are available for public testing at http://www.mat.uc.pt/dms. The problems and their dimensions are given in Table 5.1. To avoid a long presentation we do not describe their mathematical formulations, which can be found in the AMPL model files. We also provide in Table 5.1 the original references for these problems — noting, however, that in some cases the formulation coded differed from the literature due to errors, mismatches or lack of information found in the corresponding papers. 5.2. Solvers tested. We have considered in our numerical studies the following publicly available solvers for MOO without derivatives: • AMOSA (Archived MultiObjective Simulated Annealing) [5] — www.isical. ac.in/~sriparna_r/software.html; • BIMADS (BI-Objective Mesh Adaptive Direct Search) [3] tested only for problems with two objective functions — www.gerad.ca/nomad/Project/ Home.html; • Epsilon-MOEA (Epsilon MultiObjective Evolutionary Algorithm) [16] — www. iitk.ac.in/kangal/codes.shtml; • GAMULTI (Genetic Algorithms for Multiobjective, MATLAB toolbox) — www.mathworks.com; • MOPSO (MultiObjective Particle Swarm Optimization) [8] — delta.cs. cinvestav.mx/~ccoello/EMOO/EMOOsoftware.html; • NSGA-II (Nondominated Sorting Genetic Algorithm II, C version) [17] — www.iitk.ac.in/kangal/codes.shtml; • NSGA-II (MATLAB implementation by A. Seshadri) — www.mathworks. com/matlabcentral/fileexchange/10429-nsga-ii-a-multi-objective -optimization-algorithm; • PAES (Pareto Archived Evolution Strategy) [29] — dbkgroup.org/knowles/ multi. However, in order to keep the paper to a reasonable size and not to confuse the reader with excessive information, we are only reporting later (see Section 6.2) a part of the numerical tests that were performed. Besides five versions of our DMS, the selected solvers were AMOSA, BIMADS, and NSGA-II (C version), since these were the ones who exhibited the best performance in the above mentioned test set. The numerical results regarding the remaining codes can be found in http://www.mat. uc.pt/dms. 5.3. Metrics and profiles used for solver comparison. In the multiobjective case, one is interested in assessing the ability of a solver to obtain points which are Pareto optimal and to compute a highly diversified subset of the whole Pareto front. With these two goals in mind, we present in the next subsections the metrics used to assess the performance of the tested solvers. While there are other metrics in the literature, we have selected the ones presented herein due to their applicability to a large set of test problems. In particular, using a metric that considers the distance from the obtained Pareto front to the true Pareto one implies the knowledge of the latter for all the problems in the test set. In addition, presenting results for a metric that only considers a small number of test problems is meaningless. Despite not including a metric that requires the true Pareto front, we present later, and for 15

Problem BK1 [25] CL1 [6] Deb41 [15] Deb512a [15] Deb512b [15] Deb512c [15] Deb513 [15] Deb521a [15] Deb521b [15] Deb53 [15] DG01 [25] DPAM1 [25] DTLZ1 [19] DTLZ1n2 [19] DTLZ2 [19] DTLZ2n2 [19] DTLZ3 [19] DTLZ3n2 [19] DTLZ4 [19] DTLZ4n2 [19] DTLZ5 [19] DTLZ5n2 [19] DTLZ6 [19] DTLZ6n2 [19] ex005 [26] Far1 [25] FES1 [25] FES2 [25] FES3 [25] Fonseca [21] I1 [24] I2 [24] I3 [24] I4 [24]

n 2 4 2 2 2 2 2 2 2 2 1 10 7 2 12 2 12 2 12 2 12 2 22 2 2 2 10 10 10 2 8 8 8 8

m 2 2 2 2 2 2 2 2 2 2 2 2 3 2 3 2 3 2 3 2 3 2 3 2 2 2 2 3 4 2 3 3 3 3

Problem I5 [24] IKK1 [25] IM1 [25] Jin1 [48] Jin2 [48] Jin3 [48] Jin4 [48] Kursawe [31] L1ZDT4 [18] L2ZDT1 [18] L2ZDT2 [18] L2ZDT3 [18] L2ZDT4 [18] L2ZDT6 [18] L3ZDT1 [18] L3ZDT2 [18] L3ZDT3 [18] L3ZDT4 [18] L3ZDT6 [18] LE1 [25] lovison1 [33] lovison2 [33] lovison3 [33] lovison4 [33] lovison5 [33] lovison6 [33] LRS1 [25] MHHM1 [25] MHHM2 [25] MLF1 [25] MLF2 [25] MOP1 [25] MOP2 [25]

n 8 2 2 2 2 2 2 3 10 30 30 30 30 10 30 30 30 30 10 2 2 2 2 2 3 3 2 1 2 1 2 1 4

m 3 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 3 3 2 3 3 2 2 2 2

Problem MOP3 [25] MOP4 [25] MOP5 [25] MOP6 [25] MOP7 [25] OKA1 [39] OKA2 [39] QV1 [25] Sch1 [25] SK1 [25] SK2 [25] SP1 [25] SSFYY1 [25] SSFYY2 [25] TKLY1 [25] VFM1 [25] VU1 [25] VU2 [25] WFG1 [25] WFG2 [25] WFG3 [25] WFG4 [25] WFG5 [25] WFG6 [25] WFG7 [25] WFG8 [25] WFG9 [25] ZDT1 [49] ZDT2 [49] ZDT3 [49] ZDT4 [49] ZDT6 [49] ZLT1 [25]

n 2 3 2 2 2 2 3 10 1 1 4 2 2 1 4 2 2 2 8 8 8 8 8 8 8 8 8 30 30 30 10 10 10

m 2 2 3 2 3 2 2 2 2 2 2 2 2 2 2 3 2 2 3 3 3 3 3 3 3 3 3 2 2 2 2 2 3

Table 5.1 A description of our test set. Recall that n is the number of variables and m is the number of objective functions.

illustrative purposes, numerical results for some selected solvers on a small subset of problems where such information is available. 5.3.1. Performance profiles. In order to present values of the different metrics for all problems and all solvers considered, we have used the so-called performance profiles, as suggested in [20] (see also [45] and the references therein for the use of performance profiles in global derivative-free optimization). Performance profiles are depicted by a plot of a cumulative distribution function ρ(τ ) representing a performance ratio for the different solvers. Let S be the set of solvers and P be the set of problems. Let tp,s denote the performance of the solver s ∈ S on the problem p ∈ P 16

— lower values of tp,s indicate better performance. The performance ratio is defined by first setting rp,s = tp,s / min{tp,¯s : s¯ ∈ S}, for p ∈ P and s ∈ S. Then, one defines ρs (τ ) = (1/|P|)|{p ∈ P : rp,s ≤ τ }|. Thus, the value of ρs (1) is the probability of the solver s winning over the remaining ones. If we are only interested in determining which solver is the best (in the sense of winning the most), we compare the values of ρs (1) for all the solvers. At the other end, solvers with the largest probabilities ρs (τ ) for large values of τ are the most robust ones (meaning the ones that solved the largest number of problems in P). 5.3.2. Purity metric. The first metric considered by us is called Purity [4] and is used to compare the Pareto fronts obtained by different solvers. Again, let S be the set of solvers and P be the set of problems. Let Fp,s denote the approximated Pareto front determined by the solver s ∈ S for problem p ∈ P. Let also Fp denote an approximation to the true Pareto front of problem p, calculated by first forming ∪s∈S Fp,s and then removing from this set any dominated points. The Purity metric Fp consists then in computing, for solver s ∈ S and problem p ∈ P, the ratio cp,s /cp,s , Fp where cp,s = |Fp,s ∩ Fp | and cp,s = |Fp,s |. This metric is thus represented by a number Fp t¯p,s = cp,s /cp,s between zero and one. Higher values for t¯p,s indicate a better Pareto front in terms of the percentage of nondominated points. When using performance profiles to analyze the performance of the solvers measured by the Purity metric, we need to set tp,s = 1/t¯p,s (then, again, lower values of tp,s indicate better performance). Note that when a solver s is not able to obtain a single nondominated point in Fp , we obtain t¯p,s = 0, and thus tp,s = +∞, meaning that solver s was ‘unable’ to solve problem p. The Purity metric has shown to be sensitive to the number and type of solvers considered in a comparison. In fact, when two ‘similar’ solvers produce similar approximated Pareto fronts, their performance under the Purity metric deteriorates significantly since many of these points will dominate each other. This effect will then let a third solver easily win among the three. Thus, we decided to only compare solvers in pairs when using the Purity metric. Still, since we have two solvers and a large number of problems, we present the results using performance profiles. An additional difficulty is the inclusion of stochastic solvers in numerical comparisons. Since two different runs of such solvers may produce different solutions, we decided to make 10 runs for each stochastic solver on each single problem. From these 10 runs, we then selected the best and the worst run. The best run simply consists of the run that has the higher percentage of nondominated solutions when compared to the remaining ones (considering as a reference Pareto front the one obtained from the ten runs performed). In a similar way, the worst run is selected as the one with the lowest percentage of nondominated points. 5.3.3. Spread metrics. The second type of metrics used by us tries to measure the extent of the spread achieved in a computed Pareto front. Since we are interested in computing a set of points that span the entire true Pareto front, the proposed metrics have to consider first ‘extreme points’ in the objective function space Rm , which will be the same for the application of the metrics on any of the obtained fronts. The description of the computation of such ‘extreme points’ will be given later in this subsection. We considered essentially two formulae for the spread metrics. The first formula attempts at measuring the maximum size of the ‘holes’ of an approximated Pareto front. Let us assume that solver s ∈ S has computed, for problem p ∈ P, an approximated Pareto front with N points, indexed by 1, . . . , N , 17

to which we add the ‘extreme points’ mentioned above and indexed by 0 and N + 1. The metric Γ > 0 consists of setting   Γ = Γp,s = max max {δi,j } , (5.1) i∈{0,...,N }

j∈{1,...,m}

where δi,j = (fi+1,j − fi,j ) (and we assume that the objective function values have been sorted by increasing order for each objective j). For m = 2, the quantities δi,j , i = 0, . . . , N , j = 1, . . . , m, are depicted in Figure 5.1. In this case, the metric reduces to consider the maximum distance in the infinity norm between consecutive points in the approximated Pareto front.

Computed extreme points

δ

Obtained points

f

2

N,2

δN−1,2

δ

δ

N−1,1

0,2

δ0,1

δN,1

f1

Fig. 5.1. Distances between points in an approximated Pareto front to be used by the metrics Γ and ∆. For simplicity, we depict the case for m = 2.

In [17] a different metric was proposed for m = 2 to indicate how well the points are distributed in an approximated Pareto front. The major drawback of that measure is that it cannot be easily extended to problems with more than two components in the objective function. In fact, it uses the concept of consecutive points lying in the approximated Pareto front, which is not possible to define without ambiguity for m > 2. The following formula, however, extends the measure of distribution of an approximated front, for higher dimensional objective spaces (m ≥ 2): ! PN −1 δ0,j + δN,j + i=1 |δi,j − δ¯j | , (5.2) ∆ = ∆p,s = max j∈{1,...,m} δ0,j + δN,j + (N − 1)δ¯j where δ¯j , for j = 1, . . . , m, is the average of the distances δi,j , i = 1, . . . , N − 1. In our numerical experience we have nevertheless compared the proposed metric of [17] with (5.2) for m = 2, obtaining very similar results. Thus, we decided to use (5.2), which allows us to include in the test set problems with more than two components for the objective function. Regarding the computation of the ‘extreme points’ in the objective function space Rm , since the true Pareto front is not known for the majority of the problems in the test set, they must be determined, for each problem p ∈ P, from the obtained Pareto fronts Fp,s for all s ∈ S. Moreover, in an attempt to have information as good as possible, we considered all runs of all solvers (including the ones for which the results are not reported in this paper). For each problem, we first removed 18

the dominated points from the reunion of all these fronts. Then, for each component of the objective function, we selected the pair corresponding to the highest pairwise distance measured using fj (·). Note that this procedure can certainly be expensive if we have many points in all these fronts, but such computation can be implemented efficiently and, after all, it is part of the benchmarking and not of the optimization itself. We also need to use performance profiles when analyzing the results measured in terms of the Γ and ∆ metrics since, again, one has the issue of having several solvers on many problems. In these cases, we have set tp,s = Γp,s or tp,s = ∆p,s depending on the metric considered. 5.3.4. Data profiles. One possible way of assessing how well derivative-free solvers perform in terms of the number of evaluations is given by the so-called data profiles proposed in [37] for single objective optimization. Suppose that there is only one objective function f (x). For each solver, a data profile consists of a plot of the percentage of problems that are solved for a given budget of function evaluations. Let hp,s be the number of function evaluations required for solver s ∈ S to solve problem p ∈ P (up to a certain accuracy). The data profile cumulative function is then defined by ds (σ) =

1 |{p ∈ P : hp,s ≤ σ}|. |P|

(5.3)

A critical issue related to data profiles is when a problem is considered as being solved. The authors in [37] suggested that a problem is solved (up to some level ε of accuracy) when f (x0 ) − f (x) ≥ (1 − ε)(f (x0 ) − fL ),

(5.4)

where x0 is the initial guess and fL is the best obtained objective function value among all solvers. In the multiobjective case we need to consider instead a reference Pareto front Fp in order to determine whether a problem p ∈ P has been solved or not. Then, a solver s is said to solve problem p, up to an accuracy of ε, if the percentage of points obtained in the reference Pareto front Fp is equal to or greater than 1 − ε, i.e., if |Fp,s ∩ Fp | ≥ 1 − ε, |Fp |/|S|

(5.5)

where Fp,s is the approximated Pareto front obtained by solver s on problem p. Note that in (5.5) the number of points in Fp is divided by the number of solvers in S in an attempt to consider that all solvers are expected to contribute equally to the reference Pareto front. The reference Pareto front can be computed in a number of possible ways depending on the choice of solvers (and on how long we let them run). To have meaningful results for our data profiles (in other words, a significant number of points in the numerator of (5.5)), we considered only the solvers in the set S chosen for comparison and a maximum number of 5000 function evaluations. The reference Pareto front is then computed by forming the union of the output fronts of the solvers and eliminating from there all the dominated points. Following [37], we also divided σ in (5.3) by n + 1 (the number of points needed to build a simplex gradient). Finally, note also that we did not consider any spread 19

metric for data profiles since such metrics might not decrease monotonically with the budget σ of function evaluations (a consequence of this fact would be that a problem could be considered unsolved after had been considered solved earlier in the running sequence). 6. Numerical experience. 6.1. Comparing different DMS variants. The simplest possible version of direct multisearch (DMS), Algorithm 2.1, initializes the list of nondominated points with a singleton (L0 = {(x0 ; α0 )}) and considers an empty search step in all iterations. This version is referred to as DMS(1). Since no initial guess has been provided along with the majority of the problems in our test set, it was our responsibility to define a default value for the initial point x0 to be used in DMS(1). A reasonable (perhaps the most neutral) choice is x0 = (u + ℓ)/2. Since DMS is competing against population based algorithms, it is desirable to equip it with the possibility of starting from an initial list different from a singleton. Such a list can be computed by first generating a set S0 of points and then eliminating from those the dominated ones. Let S0nd denote the resulting set. The initial list is then given by L0 = {(x; α0 ), x ∈ S0nd }. We considered the three following ways of generating S0 (taking |S0 | = n and S0 ⊆ Ω = [ℓ, u] in all of them): • DMS(n,line), where S0 is formed by equally spaced points on the line connecting ℓ and u, i.e., S0 = {ℓ + (i/(n − 1))(u − ℓ), i = 0, . . . , n − 1}; • DMS(n,lhs), where S0 is generated using the Latin Hypercube Sampling strategy (see [35]). In this strategy, a multi-interval in Rn is partitioned into n multi-subintervals of equal dimension and points are uniformly randomly generated in each one of these multi-subintervals. The Latin Hypercube Sampling strategy generates random points by randomly permuting these points among the multi-subintervals. Our numerical implementation uses the MATLAB function lhsdesign from the Statistics Toolbox, followed by a shifting and scaling of the generated points in [0, 1]n to the multi-interval [ℓ, u]; • DMS(n,rand), where the n elements of S0 are uniformly randomly generated in the multi-interval [ℓ, u] (see, for instance, [40]). In this case, our numerical implementation uses the MATLAB function rand, followed by a shifting and scaling of the generated points in [0, 1]n to the multi-interval [ℓ, u]. Algorithm 2.1 allows for a variety of ways of selecting the trial list from the filtered list. We chose to work with Algorithm 2.3, meaning that Ltrial = Lf iltered . The strategy chosen to manage the list consisted of always add points to the end of the list and move a point already selected as a poll center to the end of the list (at the end of an iteration). For all the variants tested (DMS(1), DMS(n,line), DMS(n,lhs), and DMS(n,rand)), we chose1 Dk = [In − In ], where In is the identity matrix of order n. We have chosen ρ¯(·) as the constant, zero vector of dimension m. The step size parameter was halved in unsuccessful iterations and maintained in successful ones. Note that since the search step is empty these choices respect the requirements for global convergence by integer lattices (see Section A.1). Also, for all variants, we picked α0 = 1 and adopted a stopping criterion consisting 1 It is important to note that the result of Theorem 4.9 was derived under the assumption that the set of refining directions was dense in the unit sphere. We also tried in our numerical setting to use a poll set Dk equal to [Qk − Qk ] (where Qk is an orthogonal matrix computed by randomly generating the first column) but the results were not better.

20

Purity performance profile 1 0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

ρ

ρ

1 0.9

0.2

DMS(1) DMS(n,line)

0.1 0 0.5

1

1.5

2

τ

2.5

0.1

3 5

10

15 τ

20

25

0

Fig. 6.1. Comparing DMS(n,line) and DMS(1) based on performance profiles of the Purity metric. Purity performance profile with the best of 10 runs

Purity performance profile with the worst of 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

DMS(n,line) DMS(n,lhs)

0.1 0

5

10

15

20

25 τ

30

35

40

45

ρ

0.9

0 50 200 400 600 800 1000 τ

0

ρ

1

ρ

ρ

1 0.9

0.2

DMS(n,line) DMS(n,lhs) 1

2

3

4

5

6

τ

7

8

9

10

11

0.1

12

200

τ

400

0 600

Fig. 6.2. Comparing DMS(n,line) and DMS(n,lhs) based on performance profiles of the Purity metric.

of the step size αk being lower than a predefined threshold αǫ = 10−3 for all points in the list or a maximum of 20000 objective function evaluations. Figures 6.1–6.3 depict performance profiles of the Purity metric for the four above mentioned variants of DMS. When a stochastic variant is involved (DMS(n,lhs) or DMS(n,rand)), the figures show the best and worst run comparisons as explained in Subsection 5.3.2. We can easily see that DMS(n,line) is the best variant, either in terms of efficiency or robustness, although the gains when comparing to DMS(1) are not overwhelming. In fact, reading the values of the curves of Figure 6.1 for τ = 1, we can observe that both DMS(n,line) and DMS(1) are able to attain the best metric value for close to 70% of the problems. In terms of robustness, and reading the same curves but now for large values of τ , we observe that both DMS(n,line) and DMS(1) are able to provide at least one nondominated point for slightly more than 90% of the problems. However, DMS(n,line) is significantly better than DMS(n,lhs) (see Figure 6.2) and DMS(n,rand) (see Figure 6.3), in terms of both efficiency and robustness, even when considering the best Pareto front obtained for 10 runs. For the sake of brevity, we do not provide pairwise comparisons among DMS(1), DMS(n,lhs), and DMS(n,rand). The performance profiles of the spread metrics Γ and ∆ are given in Figure 6.4 for average values of the stochastic variants (the minimum and maximum values were also analyzed and do not change the conclusions stated next). In general, we can say that DMS(1) and DMS(n,line) exhibit a similar performance in terms of both metrics, better than the remaining ones regarding efficiency. 21

Purity performance profile with the best of 10 runs

Purity performance profile with the worst of 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

DMS(n,line) DMS(n,rand)

0.1 0

5

10

15

20

τ

25

30

35

40

200 300 400 500 τ

ρ

0.9

0

ρ

1

ρ

ρ

1 0.9

0.2

DMS(n,line) DMS(n,rand)

0

10

20

30

τ

40

50

60

70

0.1

800

1000

τ

1200

0 1400

Fig. 6.3. Comparing DMS(n,line) and DMS(n,rand) based on performance profiles of the Purity metric. Average Γ performance profile for 10 runs

Average ∆ performance profile for 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.1

0.1

DMS(1) DMS(n,line) DMS(n,lhs) DMS(n,rand)

0.2 0.1 0

1

2

3

4

5

τ

6

7

8

9

10

50

100 τ

150

ρ

0.9

0 200

0

ρ

1

ρ

ρ

1 0.9

0.3

DMS(1) DMS(n,line) DMS(n,lhs) DMS(n,rand) 1

1.2

1.4

τ

1.6

1.8

0.2 0.1

2 5

10

τ

15

0 20

Fig. 6.4. Comparing DMS(1), DMS(n,line), DMS(n,lhs), and DMS(n,rand) based on performance profiles of the Γ (left) and ∆ (right) metrics (taking average values for stochastic variants).

6.2. Comparing DMS to other solvers. In this section we present a comparison of the DMS(n,line) variant against the selected solvers AMOSA, BIMADS, and NSGA-II (C version). The selected solvers have been tested using their default parameters values except for the population size and number of iterations (generations). For AMOSA, we considered an initial temperature of 100, a final temperature of 2.5 × 10−6 , and a cooling factor of 0.6, yielding a total of 20650 objective function evaluations. For NSGA-II (C version), we set a population of 100 points for 200 generations, yielding a total of 20000 objective function evaluations. As mentioned before, for the DMS(n,line) solver, we imposed a stopping criterion consisting of αk < αǫ = 10−3 for all points in the list or a maximum of 20000 objective function evaluations. While AMOSA and NSGA-II (C version) always use the objective function evaluations budget, the DMS(n,line) may stop earlier due to the convergence of all the points in the list to the requested step size accuracy. For BIMADS, a limit of 20000 objective function evaluations is also imposed. The BIMADS delta criteria was set to true meaning that the runs are also stopped when the step or mesh size parameter falls below a threshold (which is set in some problem dependent way). From the performance profile of Figure 6.5, we can observe that, when using the Purity metric as a comparison measure, DMS(n,line) performs better than BIMADS in terms of efficiency, being about the same with respect to robustness. Figure 6.6 compares DMS(n,line) to AMOSA, also in terms of the Purity metric, the former being better for both the best and worst Pareto fronts obtained by AMOSA. Considering the performance profiles plotted in Figure 6.7 for the Purity metric as well, we can conclude that DMS(n,line) performs better than NSGA-II (C version) in terms of 22

Purity performance profile 1 0.9 0.8 0.7

ρ

0.6 0.5 0.4 0.3 0.2

DMS(n,line) BIMADS

0.1 0

1

1.5

2

2.5

3

3.5

τ

4

4.5

5

5.5

Fig. 6.5. Comparing DMS(n,line) and BIMADS based on performance profiles of the Purity metric (only problems with two objective functions were considered). Purity performance profile with the best of 10 runs

Purity performance profile with the worst of 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

DMS(n,line) AMOSA

0.1 0

10

20

30

40

50

60 τ

70

80

90

100

110

ρ

0.9

0 200 400 600 800 1000 τ

0

ρ

1

ρ

ρ

1 0.9

0.2

DMS(n,line) AMOSA 10

20

30

40

50

60 τ

70

80

90

100

110

0.1 0 200 400 600 800 1000 τ

Fig. 6.6. Comparing DMS(n,line) and AMOSA based on performance profiles of the Purity metric.

efficiency. Regarding robustness, and also looking at Figure 6.7, DMS(n,line) slightly outperforms NSGA-II (C version) when considering its worst Pareto front, and slightly looses compared to its best Pareto front. Figure 6.8 depicts the performance profiles using the spread metrics Γ and ∆ (see (5.1) and (5.2)) for problems where m = 2 (again we only show the results for average values of the stochastic variants as the ones for minimum and maximum values do not affect our conclusions). One can observe that DMS(n,line) exhibits the best overall performance for the Γ matric, although NSGA-II (C version) is slightly more efficient in terms of the ∆ metric. Such conclusions are true mainly in terms of efficiency, since the four solvers seem to be equally robust under both metrics. These conclusions are also supported from the performance profiles of Figure 6.9 where all the problems are considered (m ≥ 2) and BIMADS is excluded due to its limitation to m = 2. As previously mentioned, we did not use any metric which required the knowledge of the true Pareto front. This set is known, however, for some of the problems, such as Problems ZDT1–ZDT4 and ZDT6 (and is discontinuous for ZDT3). When the true Pareto front is known, which is the case for these five problems (see http://www.tik.ee.ethz.ch/sop/download/supplementary/testproblems), one can also use the Purity metric to compare the approximated Pareto fronts to the true one. Table 6.1 presents such results for the 5 problems under consideration. There are analytical expressions for the Pareto fronts of these problems, where f2 is given in terms of f1 . We considered discretized forms of these fronts by letting f1 vary in a equally spaced grid of step 10−5 . One can see, for problems ZDT1 and 23

Purity performance profile with the best of 10 runs

Purity performance profile with the worst of 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

DMS(n,line) NSGA−II (C version)

0.1 0 0.5

1

1.5

2

2.5

τ

3

3.5

4

4.5

5

20

τ

40

60

ρ

0.9

0

ρ

1

ρ

ρ

1 0.9

0.2

DMS(n,line) NSGA−II (C version)

0 0.5

1

1.5

2

2.5

τ

3

3.5

0.1

10

20 τ

30

0

Fig. 6.7. Comparing DMS(n,line) and NSGA-II (C version) based on performance profiles of the Purity metric. Average Γ performance profile for 10 runs

Average ∆ performance profile for 10 runs 1

1

0.9

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.1

0.1

DMS(n,line) BIMADS NSGA−II (C version) AMOSA

0.2 0.1 0

50

100

τ

150

200

250 500 1000 1500 2000 τ

ρ

0.9

0

0 0.5

ρ

1

ρ

ρ

1

0.3

DMS(n,line) BIMADS NSGA−II (C version) AMOSA 1

1.5

2

2.5

3

τ

3.5

4

4.5

5

5.5

0.2 0.1

6 1000

2000 τ

3000

0

Fig. 6.8. Comparing AMOSA, BIMADS, DMS(n,line), and NSGA-II (C version) based on performance profiles of the Γ (left) and ∆ (right) metrics (taking average values for stochastic variants); only problems with two objective functions were considered.

ZDT2, that at least 95% of the points in the approximated Pareto front computed by DMS(n,line) are not dominated by the true ones (up to a certain precision). BIMADS performed clearly the best for ZDT4. NSGA-II (C version) and AMOSA, on the other hand, were unable to obtain a single nondominated point for all problems. Finally, in Table 6.2 we provide the values of the spread metrics for the selected 4 solvers on these 5 problems. So far we have only reported numerical results about the quality of the approximated Pareto fronts, giving no indication on the number of evaluations of the objective functions made by the different solvers. While NSGA-II (C version) and AMOSA took all the available budget (20000 overall evaluations) for all the problems in the test set, BIMADS and the different versions of DMS managed to solve a number of problems without exhausting the budget. In Figures 6.10 and 6.11 we provide data profiles for the four solvers under consideration, AMOSA, BIMADS, DMS(n,line), and NSGA-II (C version), on the biobjective subset of our test set, corresponding to four values of accuracy ε = 0.5, 0.25, 0.1, 0.05. We chose to report only results for the best versions of the stochastic solvers AMOSA and NSGA-II (C version). So, for instance, in Figure 6.10 (left), we can conclude that if a budget of 1000 (simplex) evaluations is imposed, then both BIMADS and DMS(n,line) were able to solve around 54% of the problems in the sense of (5.5). These two solvers seem clearly the most efficient ones for budgets up to 2500 (simplex) evaluations, being BIMADS better for more accurate solutions and DMS(n,line) better for less accurate ones. 24

Average Γ performance profile for 10 runs

Average ∆ performance profile for 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

DMS(n,line) NSGA−II (C version) AMOSA

0.1 0

20

40

60

80

100

120 τ

140

160

180

200

220

500

ρ

0.9

0 1000 1500 2000 τ

0 0.5

ρ

1

ρ

ρ

1 0.9

0.2

DMS(n,line) NSGA−II (C version) AMOSA 1

1.5

2

2.5

3

τ

3.5

4

4.5

5

5.5

0.1

1000 6

1200 τ

0 1400

Fig. 6.9. Comparing AMOSA, DMS(n,line), and NSGA-II (C version) based on performance profiles of the Γ (left) and ∆ (right) metrics (taking average values for stochastic variants).

Problem DMS(n,line) BIMADS NSGA-II (C version, best) NSGA-II (C version, worst) AMOSA (best) AMOSA (worst)

ZDT1 0.974 0.126 0.000 0.000 0.000 0.000

ZDT2 0.950 0.176 0.000 0.000 0.000 0.000

ZDT3 0.804 0.083 0.000 0.000 0.000 0.000

ZDT4 0.029 0.915 0.000 0.000 0.000 0.000

ZDT6 0.992 0.682 0.000 0.000 0.000 0.000

Table 6.1 The Purity metric values (t¯p,s , see Section 5.3.2) for true Pareto front versus selected solvers.

6.3. Improving DMS performance. When solving practical derivative-free optimization problems, the major computational burden lies typically on the evaluation of the functions involved, which should then be the main focus when improving efficiency. Being DMS a direct-search method, the implementation of a cache could naturally lead to effective improvement of the code performance. We implemented a simple cache procedure by which before evaluating a point, one checks, using the infinity norm, if it has been previously evaluated. Then, objective function values will not be computed for points whose distance from a previously evaluated point is less than a tolerance (set to 10−3 ). Another important concern, now from the multiobjective aspect of the problem, is the capability of a solver to compute the entire Pareto front in an uniform way, or at least to generate a high number of evenly spread nondominated points. Having such goal in mind, we can use one of the spread metrics described before to order the current list Lk of nondominated points, before selecting a new poll center. For simplicity, we chose the metric (5.1). We have thus experimented a version of DMS where, at a given iteration, the new poll center will be the point corresponding to the highest value of Γ in the list (with ties broken by selecting the one with largest value of the step size parameter). The new version of DMS corresponding to the above described strategies was named DMS(n,line,cache,spread). In Figures 6.12 and 6.13, we report performing profiles comparing this new version of DMS and the best solvers tested, namely BIMADS and NSGA-II (C version), for the Purity metric. The corresponding data profiles can be found in Figure 6.14 for the five solvers under consideration. Figure 6.15 depicts the results obtained for the two spread metrics, Γ and ∆. 25

Problem

ZDT1

DMS(n,line) BIMADS NSGA-II (C version) AMOSA

0.044 0.040 0.053 0.240

DMS(n,line) BIMADS NSGA-II (C version) AMOSA

0.555 1.378 0.660 0.800

ZDT2 Γ 0.010 0.035 0.161 0.122 ∆ 0.492 1.389 0.788 0.795

ZDT3

ZDT4

ZDT6

0.537 0.197 0.172 0.186

0.125 0.145 7.135 0.094

3.808 1.791 1.894 1.652

1.464 1.604 1.326 1.284

0.696 1.942 0.986 0.915

1.184 1.313 0.998 1.287

Table 6.2 The Γ and ∆ metrics values for the selected solvers. (Only average values are provided for stochastic solvers.)

Data profile with the best of 10 runs (ε=0.5)

Data profile with the best of 10 runs (ε=0.25)

1

1

DMS(n,line) BIMADS NSGA−II (C version) AMOSA

0.9

0.9 0.8

0.7

0.7

0.6

0.6 d (σ)

0.5

s

s

d (σ)

0.8

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1 0

DMS(n,line) BIMADS NSGA−II (C version) AMOSA

0.1

0

500

1000

σ

1500

2000

0

2500

0

500

1000

σ

1500

2000

Fig. 6.10. Data profiles for AMOSA, BIMADS, DMS(n,line), and NSGA-II (C version) solvers (ε = 0.5 on the left and ε = 0.25 on the right).

From the observation of these profiles, we can conclude that the modifications considered, implemented in the variant DMS(n,line,cache,spread), led to an even further improving of the efficiency and robustness of the basic DMS algorithm. 7. Conclusions. In this paper we introduced, analyzed, and tested a new algorithmic approach for multiobjective optimization (MOO) without derivatives. This approach has been called direct multisearch (DMS) since it naturally generalizes direct search (of directional type) from single to multiobjective optimization. The principles of DMS are extremely simple. Instead of updating a single point per iteration, it updates an iterate list of feasible nondominated points. Iteration success is measured by changes in the iterate list. Each iteration of DMS includes provision for an optional search step. Polling is also applied, as in single objective optimization, at a selected point of the iterate list. Both steps can add points to the iterate list, forming a filtered intermediate list, and there is significant flexibility in the way a trial list is formed from this filtered list. The goal of DMS is to approximate the true (global, if possible) Pareto front, although theoretically one is only able to prove that there is a limit point in a stationary form of this front, as no aggregation or scalarization technique is incorporated in DMS. For this purpose, and to be able to state results for nonsmooth objective functions, we used in this paper the notion of a Clarke-Pareto stationary or critical 26

Data profile with the best of 10 runs (ε=0.1)

Data profile with the best of 10 runs (ε=0.05)

1

1

DMS(n,line) BIMADS NSGA−II (C version) AMOSA

0.9

0.8

0.7

0.7

0.6

0.6 d (σ)

0.5

s

s

d (σ)

0.8

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1 0

DMS(n,line) BIMADS NSGA−II (C version) AMOSA

0.9

0.1

0

500

1000

σ

1500

2000

0

2500

0

500

1000

σ

1500

2000

2500

Fig. 6.11. Data profiles for AMOSA, BIMADS, DMS(n,line), and NSGA-II (C version) solvers (ε = 0.1 on the left and ε = 0.05 on the right). Purity performance profile 1 0.9 0.8 0.7

ρ

0.6 0.5 0.4 0.3 0.2

DMS(n,line,cache,spread) BIMADS

0.1 0 0.5

1

1.5

2

2.5 τ

3

3.5

4

4.5

Fig. 6.12. Comparing DMS(n,line,cache,spread) and BIMADS based on performance profiles of the Purity metric (only problems with two objective functions were considered).

point. Our results can be further generalized for discontinuous objective functions following the steps in [47]. Our numerical experience has shown that DMS is a highly competitive technique for derivative-free MOO. Although we tested a few variants of DMS, in particular in what the initial list of nondominated points is concerned, there are a number of possible strategies which can be incorporated in the DMS framework and lead to further possible improvements. In fact, the performance of DMS is already remarkably good for the simple implementations tested on Sections 6.1 and 6.2 which do not incorporate any dissemination or spreading techniques particularly designed for the determination of the Pareto front. Such techniques could be easily fitted into DMS by means of an appropriate search step (such as a swarm search step; see [45, 46] for m = 1) or by selecting the poll centers using a spread metric (as reported in Section 6.3 where it has been shown that such procedure and the use of a cache have the potential to further improve the performance of the more basic versions). In addition, one could also study the introduction of quadratic polynomial interpolation models in DMS to possibly improve the efficiency of DMS in what concerns the search step (see [11] for what has been done in this respect in single objective optimization). One could also think of incorporating linear polynomial interpolation models (i.e., simplex gradients) to possibly improve the efficiency of an opportunistic DMS poll step (see [10, 12] for the single objective case). DMS could be parallelized in many different ways, a obvious one being the parallelization of polling. In fact, complete polling for MOO requires a total of m|Dk | function evaluations, which could be distributed among the available processors. A 27

Purity performance profile with the best of 10 runs

Purity performance profile with the worst of 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

DMS(n,line,cache,spread) NSGA−II (C version)

0.1 0 0.5

1

1.5

2 τ

2.5

3

3.5

10

τ

20

ρ

0.9

0 30

ρ

1

ρ

ρ

1 0.9

0.2

DMS(n,line,cache,spread) NSGA−II (C version)

0

2

4

6

τ

8

10

12

14

0.1

20 40 60 80 100 120 τ

0

Fig. 6.13. Comparing DMS(n,line,cache,spread) and NSGA-II (C version) based on performance profiles of the Purity metric. Data profile with the best of 10 runs (ε=0.5)

Data profile with the best of 10 runs (ε=0.05)

1

1

DMS(n,line) DMS(n,line,cache,spread) BIMADS NSGA−II (C version) AMOSA

0.9

0.8

0.7

0.7

0.6

0.6 d (σ)

0.5

s

s

d (σ)

0.8

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1 0

DMS(n,line) DMS(n,line,cache,spread) BIMADS NSGA−II (C version) AMOSA

0.9

0.1

0

500

1000

σ

1500

0

2000

0

500

1000

σ

1500

2000

Fig. 6.14. Data profiles for AMOSA, BIMADS, DMS(n,line), DMS(n,line,cache,spread), and NSGA-II (C version) solvers (ε = 0.5 on the left and ε = 0.05 on the right).

search step could also lead to various parallelization schemes. Finally, if the user of our methodology has some type of preference for an objective function (or for some of them), there are several places where such intention can be specified. In fact, there is flexibility to show preference (for some of the objective functions) in the initialization of the iterate list, in the search step, in the reordering of the list and selection of the iterate point (poll center), in the form of polling, and, finally, in the way the trial list is selected from the filtered list. Appendix A. Appendix. A.1. Globalization using integer lattices. When considering continuously differentiable functions, a finite set of directions which satisfies appropriate integrality requirements is enough to ensure convergence in single objective optimization. Generalized Pattern Search (GPS) [1, 30] makes use of such a set of directions by setting D = D. Assumption A.1. The set D of positive spanning sets is finite and the elements of D are of the form G¯ zj , j = 1, . . . , |D|, where G ∈ Rn×n is a nonsingular matrix and each z¯j is a vector in Zn . To deal with the presence of nondifferentiability, it is desirable to consider an infinite set of directions D, which should be dense (after normalization) in the unit sphere. However, if globalization is to be ensured by integer lattices, then some care must be taken when generating the set D, as it is the case in Mesh Adaptive Direct Search (MADS) [2], where generating iterates in integer lattices is guaranteed by the first requirement of the next assumption. 28

Average Γ performance profile for 10 runs

Average ∆ performance profile for 10 runs 1

1

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.3

0.3

0.3

DMS(n,line) DMS(n,line,cache,spread) BIMADS NSGA−II (C version) AMOSA

0.2 0.1 0

50

100

τ

150

200

250 500 1000 1500 2000 τ

ρ

0.9

0.2

0.2

0.1

0.1

0

0 0.5

ρ

1

ρ

ρ

1 0.9

0.4 0.3

DMS(n,line) DMS(n,line,cache,spread) BIMADS NSGA−II (C version) AMOSA 1

1.5

2

2.5

3

τ

3.5

4

4.5

5

5.5 1000

0.2 0.1

2000 τ

3000

0

Fig. 6.15. Comparing AMOSA, BIMADS, DMS(n,line), DMS(n,line,cache,spread), and NSGA-II (C version) based on performance profiles of the Γ (left) and ∆ (right) metrics (taking average values for stochastic variants); only problems with two objective functions were considered.

Assumption A.2. Let D represent a finite set of positive spanning sets satisfying Assumption A.1. The set D is so that the elements dk ∈ Dk ⊆ D satisfy the following conditions: 1. dk is a nonnegative integer combination of the columns of D. 2. The distance between xk and the point xk + αk dk tends to zero if and only if αk does: lim αk kdk k = 0 ⇐⇒ lim αk = 0,

k∈K

k∈K

for any infinite subsequence K. ¯ k = {dk /kdk k : dk ∈ Dk } are 3. The limits of all convergent subsequences of D positive spanning sets for Rn . The third requirement above is not used in the convergence theory when applied to nonsmooth objective functions, but is included for consistency with the smooth case and because it is part of the MADS original presentation [2]. Also, the strategy for updating the step size parameter must conform to some form of rationality. Assumption A.3. The step size parameter is updated as follows: Choose a rational number τ > 1, a nonnegative integer mmax ≥ 0, and a negative integer mmin ≤ −1. If the iteration is successful, the step size parameter is maintained or + increased by taking αk,new = τ m αk , with m+ ∈ {0, . . . , mmax }. Otherwise, the step − size parameter is decreased by setting αk,new = τ m αk , with m− ∈ {mmin , . . . , −1}. min max By setting β1 = τ m , β2 = τ −1 , and γ = τ m , the updating strategy described in Assumption A.3 conforms with those of Algorithm 2.1. An additional condition imposes that the search step will be conducted in a previously (implicitly defined) mesh (see Assumption A.4 below). We point out that poll points must also lie on the mesh (i.e., Pk ⊂ Mk ), but such a requirement is trivially satisfied from the definition of the mesh Mk given below. Assumption A.4. The search step in Algorithm 2.1 only evaluates points in [ |D| Mk = {x + αk Dz : z ∈ N0 }, x∈Ek

where Ek is the set of all the points evaluated by the algorithm previously to iteration k. 29

As a result of the previous assumptions, we can state the desired convergence result for the sequence of step size parameters, which was originally established by Torczon [44] in the context of pattern search and generalized by Audet and Dennis to GPS [1] and MADS [2] for single objective optimization. Theorem A.1. Let Assumption 4.1 hold. Algorithm 2.1 under one of the Assumptions A.1 or A.2 combined with Assumptions A.3–A.4 and ρ¯(·) = 0 generates a sequence of iterates satisfying lim inf αk = 0. k→+∞

Proof. In order to arrive to a contradiction, let us assume that there is a strictly positive lower bound for the step size parameter. Classical arguments, similar to the ones used by Torczon [44] and Audet and Dennis [1] for single objective optimization, allow us to conclude that all the iterates and poll points (i.e., points of the form xk + αk d, for d ∈ Dk ) generated by DMS (Algorithm 2.1) lie in an integer lattice. The intersection of a compact set with an integer lattice is finite and thus the number of points which can be added to the iterate list is finite. It remains to show that the algorithm cannot cycle among these finite number of points. If a point is removed from the iterate list, then it is because it is dominated by another point in the new iterate list. Thus, by transitivity, it can never be added again to the iterate list. At each successful iteration, at least one new point is added to the iterate list. Since the number of points which can be added is finite, the number of successful iterations must also be finite, which, according to the step size updating rules, contradicts the fact that there is a lower bound on the step size parameter. A.2. Globalization by imposing sufficient decrease. A different globalization strategy consists in using a forcing function, by considering ρ¯(·) = ρ(·) in Algorithm 2.1, imposing sufficient rather than simple decrease when accepting new iterates. The following result is relatively classic in nonlinear (single objective) optimization. Kolda, Lewis and Torczon [30] (see also [9, Section 7.7]) derive it in the context of direct-search methods of directional type, when applied to single objective optimization. We will need the following assumption (which, note, was already part of Assumption A.2). Assumption A.5. The distance between xk and the point xk + αk dk tends to zero if and only if αk does: lim αk kdk k = 0 ⇐⇒ lim αk = 0,

k∈K

k∈K

for all dk ∈ Dk and for any infinite subsequence K. Note that Assumption A.5 is a weak condition on the set of directions D. A normalized set of directions D dense in the unit sphere meets such a requirement. Theorem A.2. Let Assumption 4.1 hold. Algorithm 2.1, when ρ¯(·) is a forcing function and Assumption A.5 holds, generates a sequence of iterates satisfying lim inf αk = 0. k→+∞

Proof. Let us assume that lim inf k→+∞ αk 6= 0, meaning that there is α∗ > 0 such that αk > α∗ , for all k. From Assumption A.5, we then know that there is αd∗ > 0 such that αk kdk k > αd∗ , for all k and dk ∈ Dk . At each unsuccessful iteration k, the 30

corresponding step size parameter is reduced by at least β2 ∈ (0, 1), and thus the number of successful iterations must be infinite. Since ρ(·) is a non decreasing function, which satisfies ρ(t) > 0, for t > 0, there exists ρ∗ > 0 such that ρ(αk ) ≥ ρ(α∗ ) ≥ ρ∗ and ρ(αk kdk k) ≥ ρ(αd∗ ) ≥ ρ∗ , for all k and dk ∈ Dk , with ρ∗ = min(ρ(α∗ ), ρ(αd∗ )), taking into account what can happen in both the search and the poll steps. At each successful iteration, any new point added to the current iterate list will define a hypercube of length no smaller than ρ∗ in the set of points nondominated by those in the iterate list, where it will be later impossible to generate a new point. This and the fact that the number of successful iterations is infinite contradict Assumption 4.1. Acknowledgments. The authors would like to thank John Dennis (Rice University) and J¨org Fliege (University of Southampton) for interesting discussions on derivative-free multiobjective optimization. We are also grateful to two anonymous referees whose comments much improved the presentation of the paper. REFERENCES [1] C. Audet and J. E. Dennis Jr., Analysis of generalized pattern searches, SIAM J. Optim., 13 (2002), pp. 889–903. [2] , Mesh adaptive direct search algorithms for constrained optimization, SIAM J. Optim., 17 (2006), pp. 188–217. [3] C. Audet, G. Savard, and W. Zghal, Multiobjective optimization through a series of singleobjective formulations, SIAM J. Optim., 19 (2008), pp. 188–210. [4] S. Bandyopadhyay, S. K. Pal, and B. Aruna, Multiobjective GAs, quantative indices, and pattern classification, IEEE Transactions on Systems, Man, and Cybernetics – Part B: Cybernetics, 34 (2004), pp. 2088–2099. [5] S. Bandyopadhyay, S. Saha, U. Maulik, and K. Deb, A simulated anneling-based multiobjective optimization algorithm: AMOSA, IEEE Transactions on Evolutionary Computation, 12 (2008), pp. 269–283. [6] F. Y. Cheng and X. S. Li, Generalized center method for multiobjective engineering optimization, Engineering Optimization, 31 (1999), pp. 641–661. [7] F. H. Clarke, Optimization and Nonsmooth Analysis, John Wiley & Sons, New York, 1983. Reissued by SIAM, Philadelphia, 1990. [8] C. A. Coello Coello and M. S. Lechuga, Mopso: A proposal for multiple objective particle swarm optimization, in Congress on Evolutionary Computation (CEC’2002), vol. 2, Los Alamitos,USA, 2002, IEEE Computer Society, pp. 1051–1056. [9] A. R. Conn, K. Scheinberg, and L. N. Vicente, Introduction to Derivative-Free Optimization, MPS-SIAM Series on Optimization, SIAM, Philadelphia, 2009. ´ dio, J. E. Dennis Jr., and L. N. Vicente, Using simplex gradients of nonsmooth [10] A. L. Custo functions in direct search methods, IMA J. Numer. Anal., 28 (2008), pp. 770–784. ´ dio, H. Rocha, and L. N. Vicente, Incorporating minimum Frobenius norm [11] A. L. Custo models in direct search, Comput. Optim. Appl., 46 (2010), pp. 265–278. ´ dio and L. N. Vicente, Using sampling and simplex derivatives in pattern search [12] A. L. Custo methods, SIAM J. Optim., 18 (2007), pp. 537–555. [13] I. Das and J. E. Dennis, A closer look at drawbacks of minimizing weighted sums of objectives for pareto set generation in multicriteria optimization problems, Struct. Multidiscip. Optim., 14 (1997), pp. 63–69. [14] I. Das and J. E. Dennis Jr., Normal-boundary intersection: a new method for generating the pareto surface in nonlinear multicriteria optimization problems, SIAM J. Optim., 8 (1998), pp. 631–657. [15] K. Deb, Multi-objective genetic algorithms: Problem difficulties and construction of test problems, Evolutionary Computation, 7 (1999), pp. 205–230. 31

[16] K. Deb, M. Mohan, and S. Mishra, Towards a quick computation of well-spread paretooptimal solutions, in Evolutionary Multi-Criterion Optimization (EMO 2003), C. Fonseca et al., ed., vol. 2632 of Lecture Notes in Computer Science, 2003, pp. 222–236. [17] K. Deb, A. Pratap, S. Agarwal, and T. Meyarivan, A fast and elitist multiobjective genetic algorithm: NSGA-II, IEEE Transactions on Evolutionary Computation, 6 (2002), pp. 182– 197. [18] K. Deb, A. Sinha, and S. Kukkonen, Multi-objective test problems, linkages, and evolutionary methodologies, in Proceedings of the 8th Annual Conference on Genetic and Evolutionary Computation GECCO’06, New York, USA, 2006, ACM, pp. 1141–1148. [19] K. Deb, L. Thiele, M. Laumanns, and E. Zitzler, Scalable multi-objective optimization test problems, in Proceedings of the Congress on Evolutionary Computation (CEC’02), vol. 1, Honolulu, HI, 2002, IEEE Press, pp. 825–830. [20] E. D. Dolan and J. J. Mor´ e, Benchmarking optimization software with performance profiles, Math. Program., 91 (2002), pp. 201–213. [21] C. M. Fonseca and P. J. Fleming, Multiobjective optimization and multiple constraint handling with evolutionary algorithms – Part I: A unified formulation, IEEE Transactions on Systems, Man, and Cybernetics – Part A: Systems and Humans, 28 (1998), pp. 26–37. [22] R. Fourer, D. M. Gay, and B. W. Kernighan, A modeling language for mathematical programming, Management Sci., 36 (1990), pp. 519–554. [23] A. L. Hoffmann, A. Y. D. Siem, D. den Hertog, J. H. A. M. Kaanders, and H. Huizenga, Derivative-free generation and interpolation of convex pareto optimal IMRT plans, Phys. Med. Biol., 51 (2006), pp. 6349–6369. [24] S. Huband, L. Barone, L. While, and P. Hingston, A scalable multi-objective test problem toolkit, in Evolutionary Multi-Criterion Optimization (EMO 2005), C. A. Coello Coello et al., ed., vol. 3410 of Lecture Notes in Computer Science, Springer-Verlag, Berlin, Germany, 2005, pp. 280–295. [25] S. Huband, P. Hingston, L. Barone, and L. While, A review of multiobjective test problems and a scalable test problem toolkit, IEEE Transactions on Evolutionary Computation, 10 (2006), pp. 477–506. [26] C.-L. Hwang and A. S. Md. Masud, Multiple objective decision making, methods and applications: a state-of-the-art survey, vol. 164 of Lecture Notes in Economics and Mathematical Systems, Springer-Verlag, Berlin, Germany, 1979. [27] J. Jahn, Introduction to the Theory of Nonlinear Optimization, Springer-Verlag, Berlin, 1996. [28] D. Jones and T. Merhdad, Practical Goal Programming, Springer, Berlin, 2010. [29] J. D. Knowles and D. W. Corne, Approximating the nondominated front using the pareto archived evolution strategy, Evolutionary Computation, 8 (2000), pp. 149–172. [30] T. G. Kolda, R. M. Lewis, and V. Torczon, Optimization by direct search: New perspectives on some classical and modern methods, SIAM Rev., 45 (2003), pp. 385–482. [31] F. Kursawe, A variant of evolution strategies for vector optimization, in Parallel Problem Solving from Nature – PPSN I, H. P. Schwefel and R. M¨ anner, eds., vol. 496 of Lecture Notes in Computer Science, Springer-Verlag, Berlin, Germany, 1991, pp. 193–197. [32] G. Liuzzi, S. Lucidi, F. Parasiliti, and M. Villani, Multiobjective optimization techniques for the design of induction motors, IEEE Transactions on Magnetics, 39 (2003), pp. 1261– 1264. [33] A. Lovison, A synthetic approach to multiobjective optimization, SIAM J. Optim., (2011, to appear). [34] R. T. Marler and J. S. Arora, Survey of multi-objective optimization methods for engineering, Structural Multidisciplinary Optimization, 26 (2004), pp. 369–395. [35] M. D. McKay, R. J. Beckman, and W. J. Conover, A comparison of three methods for selecting values of input variables in the analysis of output from a computer code, Technometrics, 42 (2000), pp. 55–61. [36] K. Miettinen, Nonlinear Multiobjective Optimization, Kluwer Academic Publishers, New York, 1999. [37] J. J. Mor´ e and S. M. Wild, Benchmarking derivative-free optimization algorithms, SIAM J. Optim., 20 (2009), pp. 172–191, http://www.mcs.anl.gov/~ more/dfo. [38] J. Nocedal and S. J. Wright, Numerical Optimization, Springer-Verlag, Berlin, second ed., 2006. 32

[39] T. Okabe, Y. Jin, M. Olhofer, and B. Sendhoff, On test functions for evolutionary multiobjective optimization, in Parallel Problem Solving from Nature – PPSN VIII, X. Yao et al., ed., vol. 3242 of Lecture Notes in Computer Science, Springer-Verlag, Berlin, Germany, 2004, pp. 792–802. [40] T. J. Santner, B. J. Williams, and W. I. Notz, The Design and Analysis of Computer Experiments, Springer-Verlag, New York, 2003. [41] J. D. Schaffer, Multiple objective optimization with vector evaluated genetic algorithms, in Proceedings of the 1st International Conference on Genetic Algorithms, Hillsdale, NJ, 1985, Lawrence Erlbaum Associates, pp. 93–100. [42] R. Steuer, Multiple Criteria Optimization: Theory, Computation and Application, John Wiley & Sons, New York, 1986. [43] A. Suppapitnarm, K. A. Seffen, G. T. Parks, and P. J. Clarkson, A simulated annealing algorithm for multiobjective optimization, Engineering Optimization, 33 (2000), pp. 59–85. [44] V. Torczon, On the convergence of pattern search algorithms, SIAM J. Optim., 7 (1997), pp. 1–25. [45] A. I. F. Vaz and L. N. Vicente, A particle swarm pattern search method for bound constrained global optimization, J. Global Optim., 39 (2007), pp. 197–219. [46] , PSwarm: A hybrid solver for linearly constrained global derivative-free optimization, Optim. Methods Softw., 24 (2009), pp. 669–685. ´ dio, Analysis of direct searches for discontinuous functions, [47] L. N. Vicente and A. L. Custo Math. Program., (2011, to appear). [48] M. Olhofer Y. Jin and B. Sendhoff, Dynamic weighted aggregation for evolutionary multiobjective optimization: Why does it work and how?, in Proceedings of Genetic and Evolutionary Computation Conference GECCO’01, L. Spector et al., ed., San Francisco, CA, 2001, Morgan Kaufmann, pp. 1042–1049. [49] E. Zitzler, K. Deb, and L. Thiele, Comparison of multiobjective evolutionary algorithms: Empirical results, Evolutionary Computation, 8 (2000), pp. 173–195.

33